Skip to main content
Log in

A Group-Theoretic Approach to the Bifurcation Analysis of Spatial Cosserat-Rod Frameworks with Symmetry

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

We present a general approach to the bifurcation analysis of spatial framework structures with symmetry. While group-theoretic methods for bifurcation problems with symmetry are well known, their actual implementation in the context of geometrically exact frameworks is not straightforward. We consider spatial structures comprising assemblages of Cosserat rods; the main difficulty arises from the nonlinear configuration space, due to the presence of (cross-sectional) rotation fields. We avoid this via a single-rod formulation, developed earlier by one of the authors, whereby the governing equations are embedded in a slightly enlarged linear space. The field equations for a framework then comprise the assembly of all such rod equations, supplemented by compatibility and equilibrium conditions at the joints. We demonstrate their equivariance under the natural symmetry group action, and the implementation of group-theoretic methods is now clear within the enlarged linear space. All potential generic, symmetry-breaking bifurcations are predicted a-priori. We then employ an open-source path-following code, which can detect and compute simple, one-dimensional bifurcations; multiple bifurcation points are beyond its capabilities. For the latter, we construct symmetry-reduced problems implemented by appropriate substructures. Multiple bifurcations are rendered simple, and the path-following code is again applicable. We first analyze a simple tripod framework, providing all details of our methodology. We then treat a more complex hexagonal space frame via the same approach. Both structures exhibit simple and double bifurcation points.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig.2
Fig.3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig.9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16

Similar content being viewed by others

Data Availability

The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

References

  • A. Bossavit, Symmetry, groups, and boundary value problems. a progressive introduction to noncommutative harmonic analysis of partial differential equations in domains with geometrical symmetry,” Comput. Meth. Appl. Mech. Engrg., vol. 56, no. 2, pp. 167–215, 1986.

  • A. Vanderbauwhede, Local bifurcation and symmetry, Pitman, 1980.

  • Combescure, C., Elliott, R.S., Triantafyllidis, N.: Deformation patterns and their stability in finitely strained circular cell honeycombs. J. Mech. Phys. Solids 142, 103976 (2020)

    Article  MathSciNet  Google Scholar 

  • Combescure, C., Henry, P., Elliott, R.S.: Post-bifurcation and stability of a finitely strained hexagonal honeycomb subjected to equi-biaxial in-plane loading. Int. J. Solids Struct. 88–89, 296–318 (2016)

    Article  Google Scholar 

  • D. Sattinger, Group theoretic methods in bifurcation theory, Springer-Verlag, 1979.

  • Dellnitz, M., Werner, B.: Computational methods for bifurcation problems with symmetries - with special attention to steady state and Hopf bifurcation points. J. Comput. Appl. Math. 26, 97–123 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  • E. Doedel, R.C. Paffenroth, A.R. Champneys, T.F. Fairgrieve, Y.A. Kuznetsov, B.E. Oldeman, B. Sandstede, X. Wang, AUTO-07P: Continuation and bifurcation software for ordinary differential equations (2007) http://indy.cs.concordia.ca/auto/.

  • Elliott, R.S., Triantafyllidis, N., Shaw, J.A.: Reversible stress-induced martensitic phase transformations in a bi-atomic crystal. J. Mech. Phys. Solids 59, 216–236 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • H. Kielhöfer, Bifurcation theory, 2nd Ed., Springer, 2012

  • Healey, T.J.: A group-theoretic approach to bifurcation problems with symmetry. Comput. Meth. Appl. Mech. Engrg. 67, 257–295 (1988a)

    Article  MathSciNet  MATH  Google Scholar 

  • Healey, T.J.: Global bifurcation and continuation in the presence of symmetry with an application to solid mechanics. SIAM J. Math Anal. 19, 824–840 (1988b)

    Article  MathSciNet  MATH  Google Scholar 

  • Healey, T.J.: Material symmetry and chirality in nonlinearly elastic rods. Math. Mech. Solids 7, 405–420 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  • Healey, T.J., Mehta, P.G.: Straightforward computation of spatial equilibria of geometrically exact Cosserat rods. Int. J. Bifurcation Chaos 15, 949–965 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  • Ikeda, K., Murota, K.: Bifurcation analysis of symmetric structures using block-diagonalization. Comput. Methods Appl. Mech. Engrg. 86, 215–243 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  • J. Glowacki, Computation and visualization in multiscale modelling of DNA mechanics, doctoral dissertation, Chaire D’Analyse Appliquée, Ecole Polytechnique Federale de Lausanne, Suisse, 2016.

  • K. Ikeda, K. Murota, Imperfect bifurcation in structures and materials, Springer, 2010

  • M. Golubitsky, I. Stewart, D.G. Schaeffer, Singularities and groups in bifurcation theory, Vol. II, Springer-Verlag, 1988.

  • P. Chossat, R. Lauterbach, Methods in Equivariant Bifurcations and Dynamical Systems, World Scientific, 2000.

  • R. McWeeny, An introduction to group theory and its applications, Dover Publications, 2002

  • S.J. Mohand, Group theoretic framework for fem analysis of symmetric structures, doctoral dissertation, department of mechanical engineering, indian institute of science, Bangalore, India, 2004

  • S.J. Britvec, The Stability of Elastic Systems, Pergamom Press, 1973.

  • Simo, J.C., Vu-Cuoc, L.: A three-dimensional finite-strain rod model Part II: computational aspects. Comput. Methods Appl. Mech. Engrg. 58, 79–116 (1986)

    Article  MATH  Google Scholar 

  • Wohlever, J.C., Healey, T.J.: A group-theoretic approach to global bifurcation analysis of an axially compressed cylindrical shell. Comput. Methods Appl. Mech. Engrg. 122, 315–349 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  • Yu, T., Deier, L., Marmo, F., Gabriele, S., Parascho, S., Adriaenssens, S.: Numerical modeling of static equilibria and bifurcations in bigons and bigon rings. J. Mech. Phys. Solids 152, 104459 (2021)

    Article  MathSciNet  Google Scholar 

  • Zingoni, A.: Group-theoretic exploitations of symmetry in computational solid and structural mechanics. Int. J. Numer. Meth. Engng. 79, 253–289 (2009)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The work of TJH was supported in part by the National Science Foundation through grant DMS-2006586, which is gratefully acknowledged. CJC acknowledges support from Project ISITE FUTURE MAJOR, University Paris-Est Marne-la-Vallée.

Author information

Authors and Affiliations

Authors

Contributions

CJC, JT, and TJH participated in developing the theoretical methods. CJC carried out the numerical computations, and TJH wrote the main manuscript text. All authors reviewed the manuscript.

Corresponding author

Correspondence to Christelle J. Combescure.

Ethics declarations

Conflict of interest

All authors declare that they have no conflicts of interest to disclose.

Additional information

Communicated by Eliot Fried.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix 1

Proof of Proposition 1.1

For simplicity, we suppose that the configuration of the rod is prescribed at the two ends, viz. \({\mathbf{r}}(0),{\mathbf{R}}(0),{\mathbf{r}}(L)\) and \({\mathbf{R}}(L)\) are specified. We consider smooth but arbitrary fields \({{\varvec{\upeta}}}(x),{{\varvec{\Theta}}}(x)\) that vanish at the end points with \({{\varvec{\Theta}}}^{T} \equiv - {{\varvec{\Theta}}},\) i.e., \({{\varvec{\Theta}}}\) is a skew-symmetric field. The former is an additive admissible variation for the field \({\mathbf{r}},\) in contrast to the variation in the rotation field \({\mathbf{R}}.\) Rather, a finite perturbation of \({\mathbf{R}}(s)\) reads \(\exp [\varepsilon {{\varvec{\Theta}}}(s)]{\mathbf{R}}(s),\) and an admissible first-order variation follows from \({\mathbf{R}} \to\) \(\frac{d}{d\varepsilon }\exp [\varepsilon {{\varvec{\Theta}}}]{\mathbf{R}}|_{\varepsilon = 0} = {\varvec{\Theta}}{\mathbf{R}}.\) The first-variation condition for (1.6) then follows from the calculation

$$ \frac{d}{d\varepsilon }{\mathcal{U}}[{\mathbf{r}} + \varepsilon {{\varvec{\upeta}}},{\mathbf{R}} + \varepsilon {\varvec{\Theta}}{\mathbf{R}} + o(\varepsilon )]|_{\varepsilon = 0} = 0{\text{ for all }}{{\varvec{\upeta}}},{{\varvec{\Theta}}} $$
(4.1)

From (1.6), this entails unravelling the pertinent first-order perturbations:

$$ \begin{aligned} [{\bf R} + \varepsilon {\varvec{\Theta}}{\mathbf{R}} + o(\varepsilon )]^{T} ({\dot{\mathbf{r}}} + \varepsilon {\dot{\varvec{\eta }}}) = {\mathbf{R}}^{T} {\dot{\mathbf{r}}} + \varepsilon {\mathbf{R}}^{T} ({\dot{\varvec{\eta }}} - {\varvec{\Theta \dot{r}}}) + o(\varepsilon )\hfill \\ \, = {\mathbf{R}}^{T} {\dot{\mathbf{r}}} + \varepsilon {\mathbf{R}}^{T} ({\dot{\varvec{\eta }}} + {\dot{\mathbf{r}}} \times {{\varvec{\uptheta}}}) + o(\varepsilon ), \end{aligned} $$

where \({{\varvec{\uptheta}}} = axial({{\varvec{\Theta}}}).\) In addition,

$$ \begin{aligned} {[}{\dot{\mathbf{R}}} +\varepsilon ({\dot{\varvec{\Theta }}} {\mathbf{R}}+ {\varvec{\Theta \dot{R}}}) + o(\varepsilon ){]}[{\mathbf{R}} + \varepsilon {\varvec{\Theta R}} + o(\varepsilon )]^{T} \\ = {\mathbf{K}} + \varepsilon ({\dot{\varvec{\Theta }} + {\varvec{\Theta} }{\mathbf{{K}}}} - {\mathbf{{K}}}{\varvec{\Theta} }) + o(\varepsilon ), \\ \end{aligned} $$
$$ axial\{ {\mathbf{K}} + \varepsilon ({\dot{\varvec{\Theta }}} + {\varvec{\Theta}} {{{\rm K}}} - {\mathbf{{\rm K}}}{\varvec{\Theta }}) + o(\varepsilon )\} = {{\varvec{\upkappa}}} + \varepsilon ({\dot{\varvec{\theta }}} + {{\varvec{\uptheta}}} \times {{\varvec{\upkappa}}}) + o(\varepsilon ), $$

and then

$$ \begin{gathered} {[}{\mathbf{R}} + \varepsilon {\varvec{\Theta}} {\mathbf{R}} + o(\varepsilon ){]}^{T} [{{\varvec{\upkappa}}} + \varepsilon ({\dot{\varvec{\theta }}} + {{\varvec{\uptheta}}} \times {{\varvec{\upkappa}}}) + o(\varepsilon )] \hfill \\ \, = {\mathbf{R}}^{T} {{\varvec{\upkappa}}} + \varepsilon {\mathbf{R}}^{T} ({\dot{\varvec{\theta }}} + {{\varvec{\uptheta}}} \times {{\varvec{\upkappa}}} - {\varvec{\Theta \kappa }}) + o(\varepsilon ) \hfill \\ \, = {\mathbf{R}}^{T} {{\varvec{\upkappa}}} + \varepsilon {\mathbf{R}}^{T} {\dot{\varvec{\theta }}} + o(\varepsilon ). \hfill \\ \end{gathered} $$
(4.2)

Returning to (4.1), we now find

$$ \begin{gathered} \left\langle {\delta {\mathcal{U}}[{\mathbf{r}},{\mathbf{R}}],({{\varvec{\upeta}}},{{\varvec{\uptheta}}})} \right\rangle : = \frac{d}{d\varepsilon }{\mathcal{U}}[{\mathbf{r}} + \varepsilon {{\varvec{\upeta}}},{\mathbf{R}} + \varepsilon {\varvec{\Theta}}{\mathbf{ R}} + o(\varepsilon )]|_{\varepsilon = 0} \hfill \\ \, = \frac{d}{d\varepsilon }\int_{0}^{L} {W\left( {{\mathbf{R}}^{T} {\dot{\mathbf{r}}} + \varepsilon {\mathbf{R}}^{T} ({\dot{\varvec{\eta }}} + {\dot{\mathbf{r}}} \times {{\varvec{\uptheta}}}) + o(\varepsilon ),{\mathbf{R}}^{T} {{\varvec{\upkappa}}} + \varepsilon {\mathbf{R}}^{T} {\dot{\varvec{\theta }}} + o(\varepsilon )} \right)} dx|_{\varepsilon = 0} \hfill \\ \, = \int_{0}^{L} {\left\{ {\frac{\partial W}{{\partial{\varvec {\nu}_{i} }}}({\mathbf{R}}^{T} {\dot{\mathbf{r}}},{\mathbf{R}}^{T} {{\varvec{\upkappa}}}){\mathbf{e}}_{i} \cdot [{\mathbf{R}}^{T} ({\dot{\varvec{\eta }}} + {\dot{\mathbf{r}}} \times {{\varvec{\uptheta}}})] + \frac{\partial W}{{\partial \kappa_{i} }}({\mathbf{R}}^{T} {\dot{\mathbf{r}}},{\mathbf{R}}^{T} {{\varvec{\upkappa}}}){\mathbf{e}}_{i} \cdot {\mathbf{R}}^{T} {\dot{\varvec{\theta }}}} \right\}dx} \hfill \\ \, = \int_{0}^{L} {[{\mathbf{n}} \cdot } ({\dot{\varvec{\eta }}} + {\dot{\mathbf{r}}} \times {{\varvec{\uptheta}}}) + {\mathbf{m}} \cdot {\dot{\varvec{\theta }}}]ds = 0{\text{ for all }}{{\varvec{\upeta}}},{{\varvec{\uptheta}}}, \hfill \\ \end{gathered} $$
(4.3)

where the last line follows from the use of (1.1) and (1.5). The first-variation condition (4.3) expresses the weak form of the equilibrium equations; the classical form (1.7) follows from (4.3) via integration by parts and the usual formal arguments of the calculus of variations.\(\square\)

Proof of Proposition 2.1

We assume the end conditions (2.4), (2.5). In consonance with those, we consider smooth fields \({{\varvec{\upeta}}}^{(j)} (s),_{{}} {{\varvec{\Theta}}}^{(j)} (s)\) satisfying

$$ \begin{gathered} {{\varvec{\upeta}}}^{(j)} (L) = {\mathbf{0}},_{{}} {{\varvec{\Theta}}}^{(j)} (L) = {\mathbf{O}},j = 1,2,3, \hfill \\ {{\varvec{\upeta}}}_{A} : = {{\varvec{\upeta}}}^{(1)} (0) = {{\varvec{\upeta}}}^{(2)} (0) = {{\varvec{\upeta}}}^{(3)} (0), \hfill \\ {{\varvec{\Theta}}}^{(1)} (0) = {{\varvec{\Theta}}}^{(2)} (0) = {{\varvec{\Theta}}}^{(3)} (0), \hfill \\ \end{gathered} $$
(4.4)

which we use to construct admissible variations of \(({\mathbf{r}}^{(j)} ,{\mathbf{R}}^{(j)} ),j = 1,2,3,\) respectively, as in (4.2). We express this abstractly via

$$ \begin{gathered} {\tilde{\mathcal{U}}}[{\mathbf{u}};{\mathbf{h}},\varepsilon ]: = {\mathcal{U}}({\mathbf{r}}^{(1)} + \varepsilon {{\varvec{\upeta}}}^{(1)} ,{\mathbf{R}}^{(1)} + \varepsilon {{\varvec{\Theta}}}^{(1)} {\mathbf{R}}^{(1)} ,{\mathbf{r}}^{(2)} + \varepsilon {{\varvec{\upeta}}}^{(2)} ,{\mathbf{R}}^{(2)} + \varepsilon {{\varvec{\Theta}}}^{(2)} {\mathbf{R}}^{(2)} ,{\mathbf{r}}^{(3)} + \varepsilon {{\varvec{\upeta}}}^{(3)} , \hfill \\ \, {\mathbf{R}}^{(3)} + \varepsilon {{\varvec{\Theta}}}^{(3)} {\mathbf{R}}^{(3)} ) + o(\varepsilon ), \hfill \\ \end{gathered} $$
(4.5)

where \({\mathbf{h}}: = ({{\varvec{\upeta}}}^{(1)} ,{{\varvec{\uptheta}}}^{(1)} ,{{\varvec{\upeta}}}^{(2)} ,{{\varvec{\uptheta}}}^{(2)} ,{{\varvec{\upeta}}}^{(3)} ,{{\varvec{\uptheta}}}^{(3)} ),\) with \({{\varvec{\uptheta}}}^{(i)} = axial({{\varvec{\Theta}}}^{(i)} ),\) cf. (2.25).

Then as in (4.3), we compute the first-variation condition:

$$ \begin{gathered} \left\langle {\delta {\mathcal{V}\ominus }[\lambda ,{\mathbf{u}}],{\mathbf{h}}} \right\rangle : = \frac{d}{d\varepsilon }\left\{ {{\tilde{\mathcal{U}}}[{\mathbf{u}};{\mathbf{h}},\varepsilon )] + \lambda ({\mathbf{r}}_{A} + \varepsilon {{\varvec{\upeta}}}_{A} ) \cdot {\mathbf{i}}_{3} } \right\}|_{\varepsilon = 0} = \left\langle {\delta {\mathcal{U}\ominus }[\lambda ,{\mathbf{u}}],{\mathbf{h}}} \right\rangle + \lambda {\mathbf{i}}_{3} \cdot {{\varvec{\upeta}}}_{A} \hfill \\ \, = \sum\limits_{j = 1}^{3} {\{ \frac{d}{d\varepsilon }\int_{0}^{L} {W([{\mathbf{R}}^{(j)} ]^{T} {\dot{\mathbf{r}}}^{(j)} + \varepsilon [{\mathbf{R}}^{(j)} ]^{T} ({\dot{\varvec{\upeta }}}^{(j)} + {\dot{\mathbf{r}}}^{(j)} \times {{\varvec{\uptheta}}}^{(j)} ) + o(\varepsilon ),} } \, [{\mathbf{R}}^{(j)} ]^{T} {{\varvec{\upkappa}}}^{(j)} + \varepsilon [{\mathbf{R}}^{(j)} ]^{T} {\dot{\varvec{\uptheta }}}^{(j)} + o(\varepsilon ))dx{\text{\} }}|_{\varepsilon = 0} \hfill \\ \, \quad+ \lambda {\mathbf{i}}_{3} \cdot {{\varvec{\upeta}}}_{A} \, \hfill \\ \end{gathered} $$
$$ \begin{gathered} = \sum\limits_{j = 1}^{3} {\int_{0}^{L} {\{ \frac{\partial W}{{\partial \nu_{i} }}([{\mathbf{R}}^{(j)} ]^{T} {\dot{\mathbf{r}}}^{(j)} ,[{\mathbf{R}}^{(j)} ]^{T} {{\varvec{\upkappa}}}^{(j)} ){\mathbf{e}}_{i}^{(j)} \cdot [{\mathbf{R}}^{(j)} ]^{T} ({{\dot{\varvec{\upeta}}}}^{(j)} + {\dot{\mathbf{r}}}^{(j)} \times {{{\varvec{\uptheta}}}}^{(j)} )} } \hfill \\ \, \quad+ \frac{\partial W}{{\partial \kappa_{i} }}([{\mathbf{R}}^{(j)} ]^{T} {\dot{\mathbf{r}}}^{(j)} ,[{\mathbf{R}}^{(j)} ]^{T} {{\varvec{\upkappa}}}^{(j)} ){\mathbf{e}}_{i}^{(j)} \cdot [{\mathbf{R}}^{(j)} ]^{T} {{\dot{\varvec{\uptheta}}}}^{(j)} \} dx + \lambda {\mathbf{i}}_{3} \cdot {{\mathbf{\upeta}}}_{A} { = 0} \hfill \\ { = }\sum\limits_{j = 1}^{3} {\int_{0}^{L} {\{ {\mathbf{n}}^{(j)} } } \cdot ({{\dot{\varvec{\upeta}}}}^{(j)} + {\dot{\mathbf{r}}}^{(j)} \times {{\varvec{\uptheta}}}^{(j)} ) + {\mathbf{m}}^{(j)} \cdot \user2{{\dot{\varvec{\uptheta}}}}^{(j)} \} dx + \lambda {\mathbf{i}}_{3} \cdot {{\varvec{\upeta}}}_{A} { = 0}, \hfill \\ \end{gathered} $$
(4.6)

for all admissible fields \({{\varvec{\upeta}}}^{(j)} ,{{\varvec{\uptheta}}}^{(j)} = axial{{\varvec{\Theta}}}^{(j)} ,j = 1,2,3,\) satisfying (4.4); we used (1.1) and (1.4) in going from (4.6)2 to (4.6)3. The last line of (4.6) expresses the weak form of the equilibrium equations. Integration by parts yields, use of (4.4), and the usual formal arguments of the calculus of variations then deliver the field Eqs. (2.8).\(\hfill\square\)

Proof of Proposition 2.3:

First, we note that (2.24) is trivial for the symmetry-preserving loading term: From (4.6)1, the first variation in the loading potential is simply \(\lambda {\mathbf{i}}_{3} \cdot {{\varvec{\upeta}}}_{A} \equiv \lambda {\mathbf{Gi}}_{3} \cdot {{\varvec{\upeta}}}_{A}\) for all \({\mathbf{G}} \in C_{3v} .\) So it’s enough to show the validity of (2.24) for the internal potential energy \({\mathcal{U} }[\lambda ,{\mathbf{u}}].\) Employing (4.5) and (2.26), we compute directional derivative of the invariance conditions (2.15) (ignoring the loading term):

$$ \begin{gathered} \frac{d}{d\varepsilon }{\tilde{\mathcal{U}}}[{\varvec{T}}_{G} {\mathbf{u}};{\mathbf{h}},\varepsilon )]|_{\varepsilon = 0} = \frac{d}{d\varepsilon }{\tilde{\mathcal{U}}}[{\mathbf{u}};{\mathbf{h}},\varepsilon )]|_{\varepsilon = 0} \Rightarrow \hfill \\ \, \left\langle {\delta {\mathcal{U}}[\lambda ,{\varvec{T}}_{G} {\mathbf{u}}],\tilde{\user2{T}}_{G} {\mathbf{h}}} \right\rangle = \left\langle {\delta {\mathcal{U}}[\lambda ,{\mathbf{u}}],{\mathbf{h}}} \right\rangle , \hfill \\ \end{gathered} $$
(4.7)

for all \({\mathbf{G}} \in C_{3v}\) and for all admissible \({\mathbf{h}}.\) The modified action on the space of admissible variations in (4.7)2 arises as in (4.5)–from the conversion of skew-symmetric-matrix actions to cross products. It is straightforward, if tedious, to verify that the transformations (2.26) generate a faithful representation of \(C_{3v} ,\) i.e., the mapping \({\mathbf{G}} \mapsto \tilde{\user2{T}}_{G}\) is one-to-one, and

$$ \tilde{\user2{T}}_{G} \tilde{\user2{T}}_{H} = \tilde{\user2{T}}_{GH} ,_{{}} \tilde{\user2{T}}_{G}^{ - 1} = \tilde{\user2{T}}_{{G^{T} }} ,_{{}} \tilde{\user2{T}}_{I} = {\varvec{I}}{\text{ (identity)}}{.} $$
(4.8)

Moreover, we claim that

$$ \tilde{\user2{T}}_{G}^{*} \equiv \tilde{\user2{T}}_{G}^{ - 1} {\text{ on }}C_{3v} , $$
(4.9)

where \(\tilde{\user2{T}}_{G}^{*}\) denotes the adjoint operator of \(\tilde{\user2{T}}_{G}\) with respect to the inner-product (2.27). To see this, first note that

$$ \left\langle {\tilde{\user2{T}}_{G}^{*} \tilde{\user2{T}}_{G} {{\varvec{\upsigma}}},{\mathbf{h}}} \right\rangle = \left\langle {\tilde{\user2{T}}_{G} {{\varvec{\upsigma}}},\tilde{\user2{T}}_{G} {\mathbf{h}}} \right\rangle = \left\langle {{{\varvec{\upsigma}}},{\mathbf{h}}} \right\rangle {\text{ on }}C_{3v} , $$
(4.10)

for all smooth fields \({{\varvec{\upsigma}}},{\mathbf{h}},\) the second equality of which is readily verified via (2.26), (2.27). Thus, \(\tilde{\user2{T}}_{G}^{*}\) is a left inverse. Now let \({{\varvec{\upsigma}}} = \tilde{\user2{T}}_{G}^{*} {{\varvec{\upmu}}}\) and \({\mathbf{k}} = \tilde{\user2{T}}_{G} {\mathbf{h}}\) in the second part of (4.10), leading to \(\left\langle {\tilde{\user2{T}}_{G} \tilde{\user2{T}}_{G}^{*} {{\varvec{\upmu}}},{\mathbf{k}}} \right\rangle = \left\langle {\tilde{\user2{T}}_{G}^{*} {{\varvec{\upmu}}},{\mathbf{h}}} \right\rangle { = }\) \(\left\langle {{{\varvec{\upmu}}},{\mathbf{k}}} \right\rangle {\text{ on }}C_{3v} ,\) for all \({{\varvec{\upmu}}},{\mathbf{k}},\) i.e., \(\tilde{\user2{T}}_{G}^{*}\) is also a right inverse. Finally, we choose \({\mathbf{h}} = \tilde{\user2{T}}_{G}^{*} {\mathbf{k}}\) in (4.7)2 and employ (4.9) to deduce

$$ \begin{gathered} \left\langle {\delta {\mathcal{U} }[\lambda ,{\varvec{T}}_{G} {\mathbf{u}}],{\mathbf{k}}} \right\rangle = \left\langle {\delta {\mathcal{U} }[\lambda ,{\mathbf{u}}],\tilde{\user2{T}}_{G}^{*} {\mathbf{k}}} \right\rangle \hfill \\ \, = \left\langle {\tilde{\user2{T}}_{G} \delta {\mathcal{U} }[\lambda ,{\mathbf{u}}],{\mathbf{k}}} \right\rangle , \hfill \\ \end{gathered} $$

for all \({\mathbf{G}} \in C_{3v}\) and for all admissible \({\mathbf{k}}.\) This completes the proof.\(\square\)

Appendix 2

Here we discuss the \(C_{2v} { - }\) symmetry of the transcritical bifurcation, depicted in Fig. 13, associated with irrep \(\gamma^{(5)} .\) As noted in Sect. 3, (2.33)3 and (3.12)5 reveal that \(\gamma^{(5)}\) is associated (faithfully) with \(D_{3} { - }\) symmetry on \({\mathbb{R}}^{2} ,\) but with each element appearing twice. Hence, the maximal isotropy subgroup of the fixed-point space \(span\{ {\mathbf{i}}_{1} \}\) is actually \(\Sigma = \{ \gamma_{I}^{(5)} ,\gamma_{E}^{(5)} ,\gamma_{I}^{(5)} ,\gamma_{E}^{(5)} \} ,\) which is a representation of \(D_{2} : = \{ \gamma_{I}^{(6)} ,\gamma_{E}^{(6)} ,\gamma_{{R_{\pi } }}^{(6)} ,\gamma_{{ER_{\pi } }}^{(6)} \} \subset D_{6} .\) Accordingly, we can build a projection operator on configuration space with the symmetry of \(\Sigma\) via

$$ {\varvec{P}}_{\Sigma } = \frac{1}{4}[{\varvec{T}}_{I} + {\varvec{T}}_{E} + {\varvec{T}}_{{R_{\pi } }} + {\varvec{T}}_{{ER_{\pi } }} ], $$

where we use the []11 entries of the matrices comprising \(\Sigma\) for weights (McWeeny 2002), each of which equals 1 in this case. The operators \({\varvec{T}}_{G}\) are defined via the generators (3.8). Then \({\varvec{P}}_{\Sigma } {\varvec{u}} \in {\mathcal{X}}_{o}^{\Sigma } ,\) for any configuration \({\varvec{u}} \in {\mathcal{X}}_{o} ,\) cf. (2.31), (3.2). We now observe that

$$ \begin{gathered} {\varvec{T}}_{E} {\varvec{P}}_{\Sigma } = {\varvec{T}}_{E} [{\varvec{T}}_{I} + {\varvec{T}}_{E} + {\varvec{T}}_{{R_{\pi } }} + {\varvec{T}}_{{ER_{\pi } }} ]/4 = [{\varvec{T}}_{E} + {\varvec{T}}_{I} + {\varvec{T}}_{{ER_{\pi } }} + {\varvec{T}}_{{R_{\pi } }} ]/4 = {\varvec{P}}_{\Sigma } , \hfill \\ {\varvec{T}}_{{ER_{\pi } }} {\varvec{P}}_{\Sigma } = {\varvec{T}}_{{ER_{\pi } }} [{\varvec{T}}_{I} + {\varvec{T}}_{E} + {\varvec{T}}_{{R_{\pi } }} + {\varvec{T}}_{{ER_{\pi } }} ]/4 = [{\varvec{T}}_{{ER_{\pi } }} + {\varvec{T}}_{{R_{\pi } }} + {\varvec{T}}_{E} + {\varvec{T}}_{I} ]/4 \hfill \\ \quad\quad\quad\quad = {\varvec{P}}_{\Sigma } . \hfill \\ \end{gathered} $$
(5.1)

Above we have used the fact that \({\varvec{T}}_{G} {\varvec{T}}_{H} = {\varvec{T}}_{GH}\) for all \(G,H \in C_{6v} \cong D_{6} .\) Finally, for any \({\varvec{v}} = {\varvec{P}}_{\Sigma } {\varvec{u}} \in {\mathcal{X}}_{o}^{\Sigma } ,\) (5.1) implies \({\varvec{T}}_{{R_{\pi } }} {\varvec{v}} = \user2{v,}_{{}} {\varvec{T}}_{E} {\varvec{v}} = \user2{v,}\) and \({\varvec{T}}_{{ER_{\pi } }} {\varvec{v}} = {\varvec{v}},\) i.e., the configuration possesses \(C_{2v}\) symmetry.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Combescure, C.J., Healey, T.J. & Treacy, J. A Group-Theoretic Approach to the Bifurcation Analysis of Spatial Cosserat-Rod Frameworks with Symmetry. J Nonlinear Sci 33, 32 (2023). https://doi.org/10.1007/s00332-022-09878-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00332-022-09878-7

Keywords

Mathematics Subject Classification

Navigation